Fact-checked by Grok 2 weeks ago

Exchange operator

In , the exchange operator, often denoted as \hat{P}_{12} for two particles, is a permutation that interchanges the labels of identical particles within a multi-particle , thereby classifying the under particle exchange as a fundamental property of quantum states. This operator acts on unphysical labels in one-particle wave functions to ensure that physical solutions are proper eigenfunctions, with eigenvalues of +1 for symmetric states or -1 for antisymmetric states. The exchange operator is central to the treatment of , where the of the system must commute with it to allow common eigenfunctions that respect . For bosons, the eigenvalue +1 corresponds to fully symmetric wave functions, permitting multiple particles to occupy the same and leading to phenomena like Bose-Einstein condensation. In contrast, for fermions, the eigenvalue -1 enforces antisymmetric wave functions, which directly underpin Pauli's exclusion principle by making it impossible for two fermions to share identical quantum states, such as in atomic orbitals. Mathematically, for a two-particle system, the action of the exchange operator on a \psi(1,2) yields \hat{P}_{12} \psi(1,2) = \psi(2,1), where the labels 1 and 2 denote particle coordinates or states. This formalism extends to systems with more particles via successive pairwise exchanges and is essential for constructing proper multi-particle states in , influencing applications from molecular to . The operator's properties also connect to broader exchange interactions in , where they govern electron correlations and magnetic ordering in materials.

Fundamentals

Definition

In , the is a linear that interchanges the labels of two particles within a multi-particle , ensuring the proper treatment of particle indistinguishability. This , often denoted as \hat{P}_{ij} for particles i and j, acts on the wavefunction by the coordinates or momenta associated with those particles. For a two-particle system in position space, the action of the exchange operator is explicitly \hat{P}_{12} \psi(\mathbf{r}_1, \mathbf{r}_2) = \psi(\mathbf{r}_2, \mathbf{r}_1), where \psi is the joint wavefunction and \mathbf{r}_1, \mathbf{r}_2 are the vectors of the particles./09%3A_Indistinguishable_Particles_and_Exchange/9.02%3A_The_exchange_operator_and_Paulis_exclusion_principle) An analogous form applies in space, swapping the momentum labels while preserving the overall state structure./09%3A_Indistinguishable_Particles_and_Exchange/9.02%3A_The_exchange_operator_and_Paulis_exclusion_principle) The exchange operator represents a transposition, a basic permutation in the symmetric group of particle labelings. Permutation operators, including the exchange operator, enforce the required of the total wavefunction for identical particles: symmetric under even permutations (including the ) for bosons and antisymmetric under odd permutations (such as a single exchange) for fermions./09%3A_Indistinguishable_Particles_and_Exchange/9.02%3A_The_exchange_operator_and_Paulis_exclusion_principle) Consequently, the eigenvalues of the exchange operator are +1 for bosonic states, yielding symmetric wavefunctions, and -1 for fermionic states, yielding antisymmetric wavefunctions./09%3A_Indistinguishable_Particles_and_Exchange/9.02%3A_The_exchange_operator_and_Paulis_exclusion_principle) The historical origin of the exchange operator traces to the principle of indistinguishability of identical particles, first formalized by in 1926 during his analysis of the , where he introduced symmetric and antisymmetric combinations of states to account for their identical nature. independently advanced this framework in the same year, emphasizing antisymmetric wavefunctions for particles obeying the exclusion principle, such as s. These developments laid the foundation for handling multi-particle systems in .

Exchange Symmetry for Identical Particles

In , identical particles require that the total wavefunction of a multi-particle system exhibit definite properties under particle to account for their indistinguishability. This symmetrization postulate, introduced by Dirac, stipulates that the wavefunction must be either totally symmetric or totally antisymmetric with respect to interchange of any two identical particles, depending on the particle's intrinsic nature. For bosons, which have integer spin, the total wavefunction is symmetric under , corresponding to an exchange eigenvalue of +1; this allows multiple bosons to occupy the same , leading to phenomena like Bose-Einstein condensation. In contrast, for fermions with spin, such as electrons, the wavefunction is antisymmetric under , with an eigenvalue of -1, as required by the , which prohibits two fermions from sharing the identical . In multi-particle systems, this symmetry enforcement has profound consequences, particularly for fermions, where the wavefunction must be constructed as an antisymmetrized product of single-particle states to eliminate unphysical configurations that would violate the exclusion principle. For instance, in a two-electron system like the ground state, the spin part is a symmetric triplet (parallel spins) paired with an antisymmetric spatial wavefunction, while the antisymmetric spin (antiparallel spins) requires a symmetric spatial wavefunction to ensure overall antisymmetry. This exchange symmetry underpins quantum statistics: symmetric wavefunctions lead to Bose-Einstein statistics for bosons, while antisymmetric ones yield Fermi-Dirac statistics for fermions, governing distribution functions in . The exchange operator serves as the generator of permutations within the , dictating the allowed irreducible representations for the system's based on particle type.

Mathematical Formulation

Operator Action on Wavefunctions

The exchange operator, often denoted as \hat{P}_{ij} for particles i and j, acts on the wavefunction of a system of identical particles by interchanging their labels, enforcing the symmetry requirements dictated by quantum statistics. For a two-particle product state in the space, the action is given by \hat{P}_{ij} \left( |\psi_i\rangle \otimes |\psi_j\rangle \right) = |\psi_j\rangle \otimes |\psi_i\rangle, where |\psi_i\rangle and |\psi_j\rangle are single-particle states. This transposition is unitary and Hermitian, satisfying \hat{P}_{ij}^\dagger = \hat{P}_{ij} and \hat{P}_{ij}^2 = \hat{1}, ensuring it preserves the norm and inner products of the states. In the position representation, the action of the exchange operator on spatial wavefunctions becomes explicit through coordinate swapping. For a two-particle wavefunction \psi(\mathbf{r}_1, \mathbf{r}_2), the operator yields \hat{P}_{12} \psi(\mathbf{r}_1, \mathbf{r}_2) = \psi(\mathbf{r}_2, \mathbf{r}_1). When particles possess , the operator also interchanges spin labels, so for a state \Psi_{m_1 m_2}(\mathbf{r}_1, \mathbf{r}_2), it produces \hat{P}_{12} \Psi_{m_1 m_2}(\mathbf{r}_1, \mathbf{r}_2) = \Psi_{m_2 m_1}(\mathbf{r}_2, \mathbf{r}_1). For multi-particle systems, the exchange operator extends via the of , generating all possible label rearrangements. The complete antisymmetrizer for N fermions, essential for constructing Slater determinants, is \hat{A} = \frac{1}{N!} \sum_P (-1)^p \hat{P}, where the sum runs over all N! P, p denotes the (even or odd), and \hat{P} applies the corresponding label swap to the product wavefunction. Applying \hat{A} to an unsymmetrized product \prod_{k=1}^N \psi_k(\mathbf{r}_k) yields a fully antisymmetric , vanishing if any two orbitals are identical due to the Pauli principle. For bosons, the symmetrizer replaces (-1)^p with +1. These projectors ensure wavefunctions transform correctly under exchanges, with eigenvalues \pm 1 distinguishing fermionic and bosonic statistics. An illustrative example is the of the two-electron , where the total wavefunction must be antisymmetric under electron . The unsymmetrized product is \phi_{1s}(\mathbf{r}_1) \phi_{1s}(\mathbf{r}_2) \otimes |\alpha \beta\rangle, with \phi_{1s} the hydrogenic 1s orbital and |\alpha \beta\rangle denoting spin-up and spin-down. The operator \hat{P}_{12} maps this to \phi_{1s}(\mathbf{r}_2) \phi_{1s}(\mathbf{r}_1) \otimes |\beta \alpha\rangle. For the spin state (antisymmetric: \frac{1}{\sqrt{2}} (|\alpha \beta\rangle - |\beta \alpha\rangle)), the spatial part is symmetrized as \phi_{1s}(\mathbf{r}_1) \phi_{1s}(\mathbf{r}_2) (already symmetric for identical orbitals), yielding the total antisymmetric \frac{1}{\sqrt{2}} [\phi_{1s}(\mathbf{r}_1) \alpha(1) \phi_{1s}(\mathbf{r}_2) \beta(2) - \phi_{1s}(\mathbf{r}_1) \beta(1) \phi_{1s}(\mathbf{r}_2) \alpha(2)]. In contrast, for the triplet spin state (symmetric), the spatial part would be antisymmetrized, leading to an with nodal structure that reduces electron-electron repulsion. This demonstrates how enforces , altering probability densities compared to symmetrized forms.

Properties and Eigenvalues

The exchange operator \hat{P}, which implements particle interchange in the Hilbert space of identical particles, is both unitary and Hermitian. This follows from its definition as a permutation operator, satisfying \hat{P}^\dagger = \hat{P}^{-1} = \hat{P} and \hat{P}^2 = \hat{I}, where \hat{I} is the identity operator. The unitarity ensures preservation of the inner product under particle exchange, reflecting the indistinguishability of identical particles, while the Hermitian property guarantees real eigenvalues. Consequently, the eigenvalues of \hat{P} are restricted to \pm 1. The spectral properties of the exchange operator are tied to the requirements of particles. For bosons, the eigenstates with eigenvalue +1 span the totally symmetric , invariant under even of particle labels. For fermions, the antisymmetric corresponds to eigenvalue -1 under odd (transpositions), with the overall sign determined by the of the permutation: +1 for even permutations and -1 for odd ones. In systems with N particles and M single-particle states (M > N), these eigenspaces exhibit degeneracies given by the dimensions of the symmetric and antisymmetric representations of the S_N, such as \binom{M+N-1}{N} for the bosonic case and \binom{M}{N} for the fermionic case in higher-dimensional orbital spaces. These degeneracies highlight the group-theoretic structure underlying the operator's action. The exchange operator commutes with the Hamiltonian for systems of identical particles, [\hat{P}, \hat{H}] = 0, provided the Hamiltonian is symmetric under particle relabeling, as is required by the fundamental principles of quantum mechanics for indistinguishable particles. This commutation relation ensures that the total wavefunction can be chosen as simultaneous eigenstates of \hat{H} and \hat{P}, conserving exchange symmetry during time evolution and enabling the classification of states into bosonic or fermionic sectors. In the second-quantization formalism, the exchange operator for swapping occupations between single-particle orbitals i and j is represented using creation and annihilation operators. For fermions, it takes the form \hat{P}_{ij} = a_i^\dagger a_j a_j^\dagger a_i, incorporating anticommutation relations that introduce sign factors to enforce antisymmetry; this operator acts on Fock states by interchanging particle labels while preserving the overall fermionic statistics.

Role in Many-Body Theory

In Hartree-Fock Approximation

In the Hartree-Fock approximation, the exchange operator corrects for the classical Coulomb repulsion in the effective one-electron potential by accounting for the indistinguishability of electrons through antisymmetrization of the wavefunction. The many-electron wavefunction is constructed as a Slater determinant of orthonormal spin-orbitals to enforce antisymmetry under particle exchange, given by \Psi = \frac{1}{\sqrt{N!}} \det\left[\phi_i(\mathbf{r}_j)\right], where N is the number of electrons and \phi_i are the molecular orbitals. This form ensures that the expectation value of the incorporates exchange effects from the two-electron integrals involving permutations of electron coordinates. The derivation of the exchange term in the Fock operator proceeds from the variational minimization of the for this wavefunction, leading to a non-local operator that modifies the potential. Specifically, the exchange operator \hat{K}_k associated with an occupied orbital k acts on an orbital \phi_j(\mathbf{r}) as \hat{K}_k \phi_j(\mathbf{r}) = -\left[ \int \frac{\phi_k^*(\mathbf{r}') \phi_j(\mathbf{r}')}{|\mathbf{r} - \mathbf{r}'|} \, d\mathbf{r}' \right] \phi_k(\mathbf{r}), where the total exchange operator is \hat{K} = \sum_k \hat{K}_k with the sum over occupied orbitals k, representing the correction to the electron-electron repulsion due to quantum exchange. This term arises from the antisymmetric two-electron integrals in the energy expression, effectively reducing the repulsion for electrons of the same spin. In the Hartree-Fock equations, the exchange contributes to the self-consistent Fock operator, yielding (\hat{h} + \hat{J} - \hat{K}) \phi_i = \epsilon_i \phi_i, where \hat{h} is the one-electron core Hamiltonian, \hat{J} is the Coulomb operator, and \hat{K} is the sum of exchange operators over occupied orbitals. The orbitals \phi_i are determined iteratively until self-consistency is achieved, with the exchange ensuring the correct fermionic . Exact is incorporated in the restricted Hartree-Fock method, where spatial orbitals are doubly occupied with opposite spins, maintaining spin-restricted symmetry. In unrestricted Hartree-Fock, separate spatial orbitals for alpha and beta spins allow for spin polarization but introduce approximations, such as potential spin contamination, as the exchange is computed using spin-dependent densities without enforcing pure spin states.

In

In (DFT), the operator from wavefunction-based methods is approximated through functionals of the \rho(\mathbf{r}), enabling computationally efficient treatment of effects in many-electron s. The Kohn-Sham maps the interacting onto a non-interacting reference with the same , where the exchange-correlation energy E_{xc}[\rho] encapsulates and correlation via an unknown functional. The component E_x[\rho] arises from the and is approximated in the local density approximation (LDA) by integrating the energy per particle of a gas over local densities. This yields the explicit form E_x^{\text{LDA}}[\rho] = -\frac{3}{4} \left( \frac{3}{\pi} \right)^{1/3} \int \rho^{4/3}(\mathbf{r}) \, d\mathbf{r}, derived from the exact energy for the homogeneous gas, where the prefactor ensures dimensional consistency and scaling with . The corresponding exchange potential, which enters the Kohn-Sham equations as the functional derivative v_x(\mathbf{r}) = \frac{\delta E_x[\rho]}{\delta \rho(\mathbf{r})}, is obtained by varying the LDA functional with respect to \rho. For the uniform electron gas model, this derivative simplifies to v_x(\mathbf{r}) = -\left( \frac{3}{\pi} \right)^{1/3} \rho^{1/3}(\mathbf{r}), known as the Slater potential after its approximation in atomic calculations, though originally rooted in the Dirac expression for free-electron . This local potential captures the average field but overestimates the exact Hartree-Fock by a factor of $3/2 in the uniform limit, reflecting its semi-empirical nature for inhomogeneous systems. To improve upon pure LDA exchange, which underestimates band gaps and delocalization errors, incorporate a portion of exact Hartree-Fock exchange, which depends on orbitals rather than density alone. A prominent example is the B3LYP functional, where 20% of the exchange is from exact Hartree-Fock, combined with 80% local spin density () exchange plus a gradient correction from Becke 1988, and correlation from Lee-Yang-Parr plus Vosko-Wilk-Nusair local correlation. Specifically, E_{xc}^{\text{B3LYP}} = 0.20 E_x^{\text{HF}} + 0.80 E_x^{\text{LSD}} + 0.72 \Delta E_x^{\text{B88}} + 0.19 E_c^{\text{VWN}} + 0.81 E_c^{\text{LYP}}. This admixture enhances accuracy for and excitation energies by partially restoring the nonlocal character of exact exchange while retaining DFT's efficiency. Compared to pure Hartree-Fock, which treats exchange exactly but neglects electron correlation, DFT approximations like LDA and hybrids better account for correlation effects through the E_c[\rho] term, leading to improved binding energies and structural predictions in diverse systems. However, the density-based exchange hole in these functionals remains approximate, often failing to satisfy exact constraints like the adiabatic connection, which limits precision in strongly correlated regimes.

Applications

In Quantum Chemistry

In ab initio quantum chemistry methods, the exchange operator plays a crucial role in accurately describing molecular binding energies by incorporating the antisymmetry requirements of the electronic wavefunction, which prevents unphysical behaviors in bond dissociation processes. For instance, in the unrestricted Hartree-Fock (UHF) treatment of the H₂ molecule, the exchange contribution enables a dissociation limit close to that of two hydrogen atoms with a total energy of -1 Hartree, addressing the artificial attraction that arises in the exchange-free Hartree method due to uncompensated self-interaction of electrons with the same spin, whereas restricted HF fails to dissociate properly. This exchange effect is particularly evident along the H₂ dissociation curve, where it stabilizes the proper separation of electron densities between the nuclei, yielding a binding energy close to the experimental value when combined with correlation corrections. Computationally, evaluating the in theory involves calculating two-electron exchange integrals over Gaussian basis sets, which traditionally scale as O(N⁴) with the number of basis functions N due to the need to sum over pairwise electron interactions. To mitigate this bottleneck for larger molecules, techniques such as the approximation, also known as density fitting, are employed to approximate the elements by projecting onto an auxiliary basis set, reducing the scaling to nearly O(N³) while maintaining high accuracy for exchange energies within 1 kcal/mol. These implementations are standard in software and enable practical calculations for systems with hundreds of atoms. The operator also significantly influences computed molecular properties such as ionization potentials (IPs) and energies by modulating orbital energies and response functions. In HF theory, approximates IPs as the negative of the highest occupied energy, where the exchange term contributes to the correct ordering and magnitude, often within 1-1.5 of experiment for small molecules like . For energies, time-dependent Hartree-Fock (TDHF) incorporates exchange in the response equations, capturing local excitations accurately; for example, in , TDHF predicts the π→π* transition at about 7.8 , closely matching coupled-cluster benchmarks, with exchange ensuring proper treatment of electron-hole interactions. A notable application is in predicting in organic radicals using unrestricted Hartree-Fock (UHF), where the operator induces splitting in the molecular orbitals, favoring high- ground states. In systems like nitroxide-based diradicals, UHF splitting between α and β orbitals stabilizes ferromagnetic alignment with constants J up to 100 cm⁻¹, as demonstrated in meta-phenylene-linked radicals, guiding the design of organic magnets. This approach highlights 's role in capturing intramolecular interactions without explicit .

In Condensed Matter Physics

In condensed matter physics, the exchange operator plays a crucial role in describing interactions in solid-state systems, particularly through tight-binding models that capture effects and collective behaviors. In the , which models strongly correlated electrons on a , the arises in the large on-site repulsion limit U \gg t, where t is the hopping . The effective low-energy includes an antiferromagnetic term J \sum_{\langle i,j \rangle} \mathbf{S}_i \cdot \mathbf{S}_j, with J = 4t^2 / U, derived from second-order involving virtual hopping processes between neighboring sites that mediate spin alignment without net charge transfer. This term stabilizes magnetic order in half-filled systems, highlighting how the exchange operator enforces fermionic antisymmetry to favor over direct Coulomb repulsion. The exchange operator also influences band structure calculations in (DFT) for periodic solids. In the local density approximation (LDA), the approximate potential often underestimates band gaps in semiconductors and insulators by 30-50%, as it inadequately captures the nonlocal nature of exact , leading to delocalized orbitals and reduced energies. Corrections via the , which incorporates exact and through the screened interaction in the , yield band gaps much closer to experiment, improving predictions for and transport in materials like and . In magnetic applications, the drives itinerant in metals via splitting. The determines the instability of the paramagnetic state, requiring the product of the at the N(E_F) and the Stoner parameter I (derived from the integral) to exceed 1, i.e., I N(E_F) > 1. This leads to an splitting \Delta = I \mu, where \mu is the polarization per , shifting majority and minority s and lowering the total energy through partial filling. In transition metals like iron, this explains temperatures around 1000 K without localized moments. Exchange effects are particularly evident in transition metal oxides, where strong correlations beyond mean-field approximations explain Mott insulating behavior. In compounds like or MnO, the half-filled d-shells lead to a Mott gap due to on-site Hund's exchange and Coulomb repulsion, but simple mean-field treatments overestimate metallic character; dynamical fluctuations and multiplet effects, captured in cluster extensions of the operator, reveal charge-transfer insulation and antiferromagnetic order with Néel temperatures up to 500 K.

References

  1. [1]
    9.2: The exchange operator and Pauli's exclusion principle
    Oct 10, 2020 · We introduce the exchange operator P ^ 12 : an operator which permutes the labels of the particles. This is a rather strange operator, because ...
  2. [2]
    [PDF] exchange operator formalism for an infinite family of solvable ... - arXiv
    Jan 12, 2010 · The exchange operator formalism in polar coordinates, previously consid- ered for the Calogero-Marchioro-Wolfes problem, is generalized to a ...<|control11|><|separator|>
  3. [3]
    Exchange Operator: Physics Fundamentals & Quantum Example
    Nov 1, 2023 · The Exchange Operator determines the magnetic properties of materials in Solid State Physics. It governs the exchange interaction between electrons in a ...
  4. [4]
    Chapter 12: Identical Particles - ESM Intranet Site
    Chapter 12: Identical Particles . Note. Quantum mechanics is like Belgian endives: it is an acquired taste. - ... The exchange operator is defined as: \hat{P}_{12} ...
  5. [5]
    [PDF] Identical Particles 1 Two-Particle Systems - Physics Courses
    we define the exchange operator Pij = Pji which exchanges the coordinates of ... Kemble, The Fundamental Principles of Quantum Mechanics). Such a set of ...
  6. [6]
    The development of the concept of exchange forces in the 1930s
    Jun 20, 2024 · The onset and the development of the concept of exchange force in quantum physics are historically reconstructed, starting from Heisenberg's
  7. [7]
    On the theory of quantum mechanics - Journals
    On the theory of quantum mechanics. Paul Adrien Maurice Dirac. Google ... Published:01 October 1926https://doi.org/10.1098/rspa.1926.0133. Abstract. The ...
  8. [8]
    Blurred orbits and blurred particles: Heisenberg's 1926 helium atom
    Apr 5, 2025 · Although the indistinguishability of the electrons and indeterminacy are common explanations for the blurring of electron orbits, we argue that ...Missing: identical | Show results with:identical
  9. [9]
    [PDF] Chapter 8 - Identical Particles - MIT OpenCourseWare
    May 8, 2022 · A permutation is said to be even if it is the product of an even number of transpositions, and it is said to be odd if it is the product of an ...
  10. [10]
    [PDF] Identical Particles
    The symmetrization postulate applies to exchange of any pair of identical particles of any species. We have only defined the exchange operator in these notes.
  11. [11]
    [PDF] Introductory Quantum Chemistry Chem 570a: Lecture Notes Prof ...
    Hint: Use the integral form of Dirac's delta function: δ(p − pj) = 1. 2πh. R ... X1(r1) is defined as the exchange operator,. ˆ. X1(r1) = n. X k6=j hχ0 k(2)|.
  12. [12]
    [PDF] Systems of Identical Particles - CMU Quantum Theory Group
    Mar 21, 2011 · The permutation operators form a group under multiplication which is isomorphic to the group of permutations under composition. 1.5 Symmetric ...
  13. [13]
    Particle exchange statistics beyond fermions and bosons - Nature
    Jan 8, 2025 · It is commonly believed that there are only two types of particle exchange statistics in quantum mechanics, fermions and bosons, ...
  14. [14]
    [PDF] arXiv:quant-ph/0003118v1 24 Mar 2000
    From the properties of the exchange operators ˆP(j,k) it follows that the permutation operators admit only two common eigenstates, with the eigenvalues ǫj,k ...
  15. [15]
    [PDF] arXiv:2307.01963v3 [quant-ph] 2 Jan 2024
    Jan 2, 2024 · We are concerned with continuous-time quantum walks (CTQW) of identical particles [43–49] and ... It acts as an exchange operator on the 1-fermion ...
  16. [16]
  17. [17]
    Perspective: Fifty years of density-functional theory in chemical physics
    Apr 1, 2014 · EX is the exchange energy of the Slater determinant of the Kohn-Sham orbitals, and EC is everything else. Because the correlation hole ...
  18. [18]
    [PDF] Describing static correlation in bond dissociation by Kohn–Sham ...
    Static correlation, a left-right correlation, is important in bond dissociation. Standard DFT methods underestimate this, and exchange functionals can mimic it.
  19. [19]
    Nonempirical Adiabatic Connection Correlation Functional from ...
    Mar 26, 2025 · The dissociation curve of H2 within a spin-restricted formalism ... (GGAs) for exchange and correlation with Hartree-Fock (HF) exchange ...
  20. [20]
    Resolution-of-the-identity approximation for complex-scaled basis ...
    Sep 18, 2019 · ... Hermitian RI-HF and RI-MP2 methods. In RI-HF, the Coulomb and exchange contributions to an element Fμν of the Fock matrix are approximated as.
  21. [21]
    RI-MP2: optimized auxiliary basis sets and demonstration of efficiency
    Efficient, approximate and parallel Hartree-Fock and hybrid DFT calculations. A 'chain-of-spheres' algorithm for the Hartree-Fock exchange. 2009, Chemical ...
  22. [22]
    Influence of the exchange-correlation potential in methods based on ...
    Oct 24, 2013 · Accurate exchange-correlation potentials are found to be essential for getting accurate energies. Methods employing in the Kohn-Sham self-consistency process
  23. [23]
    [PDF] Excited states from time-dependent Hartree-Fock and density ...
    Feb 9, 2015 · A fully numerical method for the time-dependent Hartree-Fock and density functional theory (TD-HF/DFT) with the Tamm-.
  24. [24]
    Effect of Hartree-Fock exact exchange on intramolecular magnetic ...
    Jan 13, 2015 · Among them, organic diradicals are the simplest magnetic system containing only two magnetic centers, which are suitable systems to investigate ...
  25. [25]
    High-Spin S = 2 Ground State Aminyl Tetraradicals - ACS Publications
    The quintet ground state with this ΔETQ corresponds to a ferromagnetic exchange coupling within the tetraradical that is more than 3 orders of magnitude ...Epr Spectroscopy · Dimers Of S = 2... · Decay Kinetics Of Aminyl...
  26. [26]
    LDA, GW, and exact-exchange Kohn-Sham scheme calculations of ...
    Nov 29, 2001 · The GW approximation is applied perturbatively either on top of the LDA, or EXX band structures in the zero order. It is argued that such a ...
  27. [27]
    Combining GW calculations with exact-exchange density-functional ...
    Feb 16, 2005 · Combining GW calculations with exact-exchange density-functional theory: An analysis of valence-band photoemission for compound semiconductors.
  28. [28]
    Collective electron ferromagnetism | Proceedings of the Royal ...
    The advantages of a collective electron treatment for ferromagnetism were pointed out some years ago in a paper (Stoner 1933) in which it was shown that such a ...
  29. [29]
    Another mechanism for the insulator-metal transition observed in ...
    Apr 10, 2008 · Mott 1 proposed the insulator model for transition metal oxides with partially filled d band. Combined with the Hubbard theory, this model ...
  30. [30]
    Doping a Mott insulator: Physics of high-temperature superconductivity
    Jan 6, 2006 · Therefore we have to go beyond the mean-field theory and take into account the fluctuation around it. In other words, the local gauge ...